Next Article in Journal
Structural, Morphological and Mechanical Properties of Concrete Slab in Traditional Buildings in Casablanca, Morocco
Previous Article in Journal
Initial Assessment of Separation Train Design and Utilities Consumption for Cyclopentyl Methyl Ether Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Catalytic Performance of Doped Ni2P Surfaces for Ammonia Synthesis

by
Abdulrahman Almithn
Department of Chemical Engineering, College of Engineering, King Faisal University, Al Ahsa 31982, Saudi Arabia
Presented at the 4th International Electronic Conference on Applied Sciences, 27 October–10 November 2023; Available online: https://asec2023.sciforum.net/.
Eng. Proc. 2023, 56(1), 53; https://doi.org/10.3390/ASEC2023-15319
Published: 21 November 2023
(This article belongs to the Proceedings of The 4th International Electronic Conference on Applied Sciences)

Abstract

:
Ammonia is a key ingredient in fertilizer production, but its synthesis using the conventional Haber–Bosch process over metal-based catalysts is energy intensive. Prior investigations revealed that metal catalysts suffer from a trade-off between N2 activation and N* binding strength, hindering their overall reactivity. Metal phosphide catalysts are promising alternatives to conventional metal catalysts due to their unique reactivity and stability. Here, we used DFT to study the catalytic performance of Ni2P catalysts doped with Fe and Ru for ammonia synthesis. We show that H-assisted N–N activation may provide a new route to circumvent the N2 dissociation scaling relationships.

1. Introduction

Ammonia (NH3) holds a crucial position in various aspects of modern society. Its significance spans from agriculture to industry, making it an essential molecule for the global economy and sustainability efforts. It serves as a fundamental building block for the production of fertilizers, which are essential for enhancing crop yields and ensuring food security for a growing global population [1]. Additionally, ammonia finds applications in diverse sectors, including chemicals, pharmaceuticals, and even renewable energy storage. The Haber–Bosch process revolutionized agricultural productivity in the early 20th century by converting atmospheric nitrogen into ammonia over Fe-based catalysts and contributing to the large-scale production of synthetic fertilizers [2]. However, despite its immense importance, ammonia synthesis presents a range of formidable challenges.
The nitrogen (N2) dissociation step is believed to be the rate-limiting step in ammonia synthesis [3,4,5,6]. Cleaving the strong N≡N bond requires high temperatures and pressures over the conventional ammonia synthesis catalysts (Fe and Ru). The surface intermediates (NHx; x = 0–3) formed during ammonia synthesis often bind too strongly to most transition metal catalysts under ambient conditions; thus, elevated temperatures and pressures are necessary to maintain a favorable exergonic nature of the overall reaction [7]. The catalytic reduction of dinitrogen to ammonia under mild conditions is a long-standing challenge in catalysis. The goal is to find a catalyst that can bind the NHx species weakly while also maintaining a low energy barrier for N2 dissociation. Computational assessments using density functional theory (DFT), however, have revealed a significant limitation: the N–N transition state energy cannot be adjusted without influencing the binding energy of NHx to the surface due to the presence of a Brønsted–Evans–Polanyi (BEP) scaling relationship for transition metal surfaces [8,9,10,11]. Decreasing the N2 activation barrier also increases the binding strength of NHx, thus hindering subsequent hydrogenation reactions and the production of NH3. These studies have concluded that the turnover rate for ammonia synthesis as a function of N* binding energy for several transition metals follows a volcano-shaped curve, with Fe and Ru being the only metals that are close to the optimum.
A logical approach to catalyst design was suggested, involving the combination of metals positioned on opposite sides of the volcano to achieve the desired interaction with nitrogen intermediates. For example, combining Co, which binds N* too weakly, with Mo, which binds N* too strongly in the Co–Mo catalyst, leads to a material that has ammonia synthesis activity similar to that of Ru [12]. Many subsequent studies have focused on finding ways to circumvent the scaling relationship using different catalytic materials, such as metal nitrides [13], hydrides [14], and oxides [15]. Electrochemical N2 reduction has also emerged as an alternative approach for synthesizing NH3 at ambient conditions, but the development of efficient electrochemical catalysts remains a key challenge [16,17,18]. Here, we explore the catalytic performance of the Ni2P catalyst doped with Fe and Ru for ammonia synthesis using DFT. We have previously shown that incorporating phosphorus atoms in nickel phosphide catalysts can alter the catalytic behavior of Ni atoms, leading to improved selectivity for various catalytic reactions [19,20,21]. The P atoms introduce weak binding surface sites that disrupt the metal ensembles. We also explore H-assisted N–N activation pathways involving diazene (N2H2) and hydrazine (N2H4) to weaken the N–N bond prior to activation.

2. Computational Methods

Periodic plane-wave DFT calculations were performed using the Vienna ab initio simulation package (VASP) [22,23,24,25] and the computational catalysis interface (CCI) [26]. The RPBE form of the generalized gradient approximation and PAW pseudopotentials were used [27,28,29,30,31]. The Ni2P(001) surface was constructed from bulk Ni2P, as described in more detail in our previous work [19]. Figure 1 shows the top and side views of the 2 × 2 Ni2P(001) surface used in this study. The bottom two atomic layers were constrained during geometric relaxation. Electronic energies were converged to within 10−6 eV, and forces were converged to less than 0.05 eVÅ−1 with a k-point mesh of 3 × 3 × 1 [32,33]. Transition state structures were identified using the nudged elastic band (NEB) method and the dimer method [34,35,36]. The Ni2P(001) surface was modified by substituting a surface Ni atom (highlighted in Figure 1a) with either a Fe or Ru atom. This procedure mirrors the approach employed in a prior investigation using doped Ni2P surfaces [37]. The formation energies of doped Ni2P surfaces were calculated using the following formula:
E f = E M / N i 2 P + E N i E N i 2 P E M
where ENi and EM are the energies of a single Ni atom and a single metal atom. The Ru–Ni2P surface has a formation energy of −2.050 eV compared to −1.129 eV for Fe–Ni2P, indicating that Ru–Ni2P is more stable. Additional calculations were run using the pure 3 × 3 Fe(110) and Ru(001) surfaces for comparison. Spin-polarized calculations were exclusively applied to the Fe(110) and Fe–Ni2P doped surfaces, as the energy discrepancy between spin-polarized and non-spin-polarized calculations exceeded 10−3 eV.

3. Results and Discussion

3.1. NHx Species Binding Energies

In this section, we investigate the preferred adsorption configuration and binding energies of different NHx species over Ni2P, as well as doped Fe–Ni2P and Ru–Ni2P catalysts. For the sake of comparison, we also included the binding energies over pure Fe(110) and Ru(001) surfaces. The binding energies shown in Table 1 are calculated with respect to the gas-phase species and bare surfaces (i.e., ΔEads = Especies/surfEspecies(g)Esurf). Diatomic nitrogen (N2) vertically adsorbs on the metal atop site (M1) over Ni2P with a binding energy of −25 kJ mol−1. This binding energy increases to −35 kJ mol−1 for Fe–Ni2P and −50 kJ mol−1 for Ru–Ni2P. For the NHx* species, the binding strength increases as the number of hydrogen atoms bonded to the N atom decreases, i.e., NH3* binds most weakly to the surface, whereas N* has the highest binding energy. This reflects the increasing stability of the more saturated NHx* species in the gas phase, and they coordinate with fewer metal atoms upon adsorption. For example, N* and NH* bind to the metal 3-fold (M3) site, whereas NH2 and NH3 bind to the bridging (M2) and atop (M1) sites, respectively (Figure 2).
We also observe a general trend in binding energies across examined surfaces where the binding strength for each species generally increases as Ni2P → Fe–Ni2P → Ru–Ni2P → Ru → Fe with few exceptions. If we consider the N* binding energy as a proxy for the catalytic activity towards N–N bond activation within N2, following the principles of Brønsted–Evans–Polanyi (BEP) scaling relations [8,9,10,11], it can be inferred that Ni2P is likely to exhibit weak activity. Conversely, Fe and Ru-doped Ni2P catalysts are expected to demonstrate higher activity compared to pure Ni2P, although falling short of the reactivity of Fe- and Ru-based catalysts. However, this also means that Fe- and Ru-based catalysts are more prone to deactivation due to the strongly bound N* or NH* than phosphide-based catalysts.

3.2. N–N Activation Reactions

Here, we examine the N–N bond cleavage reactions in N2*, N2H2*, and N2H4*, as shown in Figure 3. Over the Ru(001) surface, adsorbed N2* undergoes N–N bond activation with a forward activation barrier (ΔEa) of 168 kJ mol−1 to form two nitrogen atoms (2N*) both adsorbed on separate 3-fold sites. This reaction is slightly endothermic with a reaction energy (ΔErxn) of 13 kJ mol−1. These results are consistent with previous studies on Ru(001) flat surfaces [38]. On both Fe-Ni2P and Ru-Ni2P surfaces, N2* dissociates to 2N* with a much larger activation barrier of 350 kJ mol−1 and a highly endothermic reaction energy. This reaction starts with N2* vertically adsorbed on the doped metal atom. After the N–N bond cleavage, the N* atoms bind to nearby 3-fold sites. We have attempted this reaction and even the other H-assisted N–N bond activation reactions in N2H2 and N2H4 over the undoped Ni2P surface, but we were not able to identify stable transition states. The trends in N* binding energy discussed in Section 3.1 indeed correlate with the calculated N–N bond activation barriers over these materials.
We next examine the N–N bond activation reactions in diazine (N2H2) and hydrazine (N2H4) to investigate whether hydrogenating N2 can weaken the N–N bond prior to its activation. Figure 3 shows that the activation barrier of the N–N bond in adsorbed N2H2* dramatically decreased by more than 220 kJ mol−1 over both Fe–Ni2P and Ru–Ni2P surfaces (121 and 129 kJ mol−1, respectively) compared to direct N2* dissociation (350 kJ mol−1). Breaking the N–N bond in N2H4 also exhibits a similar barrier of 128 kJ mol−1 over Ru-Ni2P, but the barrier for this reaction over Fe-Ni2P increases to 143 kJ mol−1. We expect a similar decrease in the activation barrier of H-assisted N–N activation over a pure Ru(001) surface; however, the resulting NH* also strongly binds to the surface (Table 1), which may render the active sites inert. The more facile activation of the N–N bond in N2H2* over both Fe–Ni2P and Ru-Ni2P-doped surfaces, coupled with the weaker binding of the product NH*, may open the door for a promising pathway that avoids the current paradigm of direct N2 dissociation. However, it is important to note that further analysis of the hydrogenation reactions to produce N2H2 and eventually NH3 over these materials is necessary to fully ascertain the viability of this pathway.

4. Conclusions

In this work, we investigated the binding energies of various NHx species and the N–N bond activation energy over Ni2P catalysts doped with Fe and Ru using DFT. Our results indicate that the NHx species with fewer hydrogen atoms bind more strongly to the catalytic surfaces, and the binding strength generally increases as Ni2P → Fe–Ni2P → Ru–Ni2P → Ru → Fe. This trend is reflected in the relative activity of these catalytic surfaces towards direct N2* dissociation, with Ru being the most active and Ni2P the least active surface. However, we show that hydrogenating the N2* prior to N–N bond cleavage can dramatically decrease the activation barrier over doped Ni2P surfaces. These findings suggest that the H-assisted N–N bond activation may play a crucial role in reducing the activation energy required for ammonia synthesis over these surfaces without a concomitant increase in the binding strength of the NHx species. Further investigation into the underlying mechanisms of this catalytic process could potentially lead to enhanced strategies for nitrogen activation and ammonia synthesis.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Erisman, J.W.; Sutton, M.A.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a Century of Ammonia Synthesis Changed the World. Nat. Geosci. 2008, 1, 636–639. [Google Scholar] [CrossRef]
  2. Smil, V. Enriching the Earth: Fritz Haber, Carl Bosch, and the Transformation of World Food Production; MIT Press: Cambridge, MA, USA, 2004; ISBN 0262693135. [Google Scholar]
  3. Boudart, M. Kinetics and Mechanism of Ammonia Synthesis. Catal. Rev. 1981, 23, 1–15. [Google Scholar] [CrossRef]
  4. Honkala, K.; Hellman, A.; Remediakis, I.N.; Logadottir, A.; Carlsson, A.; Dahl, S.; Christensen, C.H.; Nørskov, J.K. Ammonia Synthesis from First-Principles Calculations. Science 2005, 307, 555–558. [Google Scholar] [CrossRef]
  5. Schlögl, R. Catalytic Synthesis of Ammonia—A “Never-Ending Story”? Angew. Chem. Int. Ed. 2003, 42, 2004–2008. [Google Scholar] [CrossRef]
  6. Ertl, G. Surface Science and Catalysis—Studies on the Mechanism of Ammonia Synthesis: The P. H. Emmett Award Address. Catal. Rev. 1980, 21, 201–223. [Google Scholar] [CrossRef]
  7. Vojvodic, A.; Medford, A.J.; Studt, F.; Abild-Pedersen, F.; Khan, T.S.; Bligaard, T.; Nørskov, J.K. Exploring the Limits: A Low-Pressure, Low-Temperature Haber–Bosch Process. Chem. Phys. Lett. 2014, 598, 108–112. [Google Scholar] [CrossRef]
  8. Logadottir, A.; Rod, T.H.; Nørskov, J.K.; Hammer, B.; Dahl, S.; Jacobsen, C.J.H. The Brønsted-Evans-Polanyi Relation and the Volcano Plot for Ammonia Synthesis over Transition Metal Catalysts. J. Catal. 2001, 197, 229–231. [Google Scholar] [CrossRef]
  9. Munter, T.R.; Bligaard, T.; Christensen, C.H.; Nørskov, J.K. BEP Relations for N2 Dissociation over Stepped Transition Metal and Alloy Surfaces. Phys. Chem. Chem. Phys. 2008, 10, 5202. [Google Scholar] [CrossRef]
  10. Nørskov, J.K.; Bligaard, T.; Hvolbæk, B.; Abild-Pedersen, F.; Chorkendorff, I.; Christensen, C.H. The Nature of the Active Site in Heterogeneous Metal Catalysis. Chem. Soc. Rev. 2008, 37, 2163. [Google Scholar] [CrossRef]
  11. Medford, A.J.; Vojvodic, A.; Hummelshøj, J.S.; Voss, J.; Abild-Pedersen, F.; Studt, F.; Bligaard, T.; Nilsson, A.; Nørskov, J.K. From the Sabatier Principle to a Predictive Theory of Transition-Metal Heterogeneous Catalysis. J. Catal. 2015, 328, 36–42. [Google Scholar] [CrossRef]
  12. Jacobsen, C.J.H.; Dahl, S.; Clausen, B.G.S.; Bahn, S.; Logadottir, A.; Nørskov, J.K. Catalyst Design by Interpolation in the Periodic Table: Bimetallic Ammonia Synthesis Catalysts. J. Am. Chem. Soc. 2001, 123, 8404–8405. [Google Scholar] [CrossRef]
  13. Zeinalipour-Yazdi, C.D.; Hargreaves, J.S.J.; Catlow, C.R.A. Nitrogen Activation in a Mars–van Krevelen Mechanism for Ammonia Synthesis on Co3Mo3N. J. Phys. Chem. C 2015, 119, 28368–28376. [Google Scholar] [CrossRef]
  14. Wang, P.; Chang, F.; Gao, W.; Guo, J.; Wu, G.; He, T.; Chen, P. Breaking Scaling Relations to Achieve Low-Temperature Ammonia Synthesis through LiH-Mediated Nitrogen Transfer and Hydrogenation. Nat. Chem. 2017, 9, 64–70. [Google Scholar] [CrossRef]
  15. Vojvodic, A.; Calle-Vallejo, F.; Guo, W.; Wang, S.; Toftelund, A.; Studt, F.; Martínez, J.I.; Shen, J.; Man, I.C.; Rossmeisl, J.; et al. On the Behavior of Brønsted-Evans-Polanyi Relations for Transition Metal Oxides. J. Chem. Phys. 2011, 134, 244509. [Google Scholar] [CrossRef]
  16. Choi, C.; Back, S.; Kim, N.-Y.; Lim, J.; Kim, Y.-H.; Jung, Y. Suppression of Hydrogen Evolution Reaction in Electrochemical N2 Reduction Using Single-Atom Catalysts: A Computational Guideline. ACS Catal. 2018, 8, 7517–7525. [Google Scholar] [CrossRef]
  17. Liu, C.; Li, S.; Li, Z.; Zhang, L.; Chen, H.; Zhao, D.; Sun, S.; Luo, Y.; Alshehri, A.A.; Hamdy, M.S.; et al. Ambient N2-to-NH3 Fixation over a CeO2 Nanoparticle Decorated Three-Dimensional Carbon Skeleton. Sustain. Energy Fuels 2022, 6, 3344–3348. [Google Scholar] [CrossRef]
  18. Wang, F.; Zhang, L.; Wang, T.; Zhang, F.; Liu, Q.; Zhao, H.; Zheng, B.; Du, J.; Sun, X. In Situ Derived Bi Nanoparticles Confined in Carbon Rods as an Efficient Electrocatalyst for Ambient N2 Reduction to NH3. Inorg. Chem. 2021, 60, 7584–7589. [Google Scholar] [CrossRef]
  19. Witzke, M.E.; Almithn, A.; Conrad, C.L.; Hibbitts, D.D.; Flaherty, D.W. Mechanisms and Active Sites for C–O Bond Rupture within 2-Methyltetrahydrofuran over Ni, Ni12P5, and Ni2P Catalysts. ACS Catal. 2018, 8, 7141–7157. [Google Scholar] [CrossRef]
  20. Almithn, A.; Alhulaybi, Z. A Mechanistic Study of Methanol Steam Reforming on Ni2P Catalyst. Catalysts 2022, 12, 1174. [Google Scholar] [CrossRef]
  21. Almithn, A.; Alghanim, S.N.; Mohammed, A.A.; Alghawinim, A.K.; Alomaireen, M.A.; Alhulaybi, Z.; Hossain, S.S. Methane Activation and Coupling Pathways on Ni2P Catalyst. Catalysts 2023, 13, 531. [Google Scholar] [CrossRef]
  22. Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals. Phys. Rev. B 1993, 47, 558–561. [Google Scholar] [CrossRef]
  23. Kresse, G.; Hafner, J. Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal–Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B 1994, 49, 14251–14269. [Google Scholar] [CrossRef]
  24. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  25. Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  26. Kravchenko, P.; Plaisance, C.; Hibbitts, D. A New Computational Interface for Catalysis. ChemRxiv, 2019; preprint. [Google Scholar]
  27. Hammer, B.; Hansen, L.B.; Nørskov, J.K. Improved Adsorption Energetics within Density-Functional Theory Using Revised Perdew-Burke-Ernzerhof Functionals. Phys. Rev. B 1999, 59, 7413–7421. [Google Scholar] [CrossRef]
  28. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  29. Zhang, Y.; Yang, W. Comment on “Generalized Gradient Approximation Made Simple”. Phys. Rev. Lett. 1998, 80, 890. [Google Scholar] [CrossRef]
  30. Blöchl, P.E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  31. Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  32. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  33. Pack, J.D.; Monkhorst, H.J. “Special Points for Brillouin-Zone Integrations”—A Reply. Phys. Rev. B 1977, 16, 1748–1749. [Google Scholar] [CrossRef]
  34. Henkelman, G.; Jónsson, H. Improved Tangent Estimate in the Nudged Elastic Band Method for Finding Minimum Energy Paths and Saddle Points. J. Chem. Phys. 2000, 113, 9978–9985. [Google Scholar] [CrossRef]
  35. Henkelman, G.; Jónsson, H. A Dimer Method for Finding Saddle Points on High Dimensional Potential Surfaces Using Only First Derivatives. J. Chem. Phys. 1999, 111, 7010–7022. [Google Scholar] [CrossRef]
  36. Jónsson, H.; Mills, G.; Jacobsen, K.W. Nudged Elastic Band Method for Finding Minimum Energy Paths of Transitions. In Classical and Quantum Dynamics in Condensed Phase Simulations; World Scientific: Singapore, 1998; pp. 385–404. [Google Scholar]
  37. Wang, G.; Shi, Y.; Mei, J.; Xiao, C.; Hu, D.; Chi, K.; Gao, S.; Duan, A.; Zheng, P. DFT Insights into Hydrogen Activation on the Doping Ni2P Surfaces under the Hydrodesulfurization Condition. Appl. Surf. Sci. 2021, 538, 148160. [Google Scholar] [CrossRef]
  38. Dahl, S.; Logadottir, A.; Egeberg, R.C.; Larsen, J.H.; Chorkendorff, I.; Törnqvist, E.; Nørskov, J.K. Role of Steps in N2 Activation on Ru(0001). Phys. Rev. Lett. 1999, 83, 1814–1817. [Google Scholar] [CrossRef]
Figure 1. The top (a) and side (b) views of the doped Ni2P(001) surface model. Labels indicate different surface sites.
Figure 1. The top (a) and side (b) views of the doped Ni2P(001) surface model. Labels indicate different surface sites.
Engproc 56 00053 g001
Figure 2. Adsorbed NHx (x = 0–3) species over Fe–Ni2P (ad) and Ru–Ni2P (eh) catalysts. The adsorption configuration and binding energies (ΔEads in kJ mol−1) are shown beneath each image.
Figure 2. Adsorbed NHx (x = 0–3) species over Fe–Ni2P (ad) and Ru–Ni2P (eh) catalysts. The adsorption configuration and binding energies (ΔEads in kJ mol−1) are shown beneath each image.
Engproc 56 00053 g002
Figure 3. Reaction coordinate diagram for N–N bond activation in N2*, N2H2*, and N2H4* over Fe–Ni2P (red), Ru–Ni2P (blue), and Ru (black). Images display the reactants, transition state, and product structures over Fe–Ni2P (top) and Ru–Ni2P (bottom).
Figure 3. Reaction coordinate diagram for N–N bond activation in N2*, N2H2*, and N2H4* over Fe–Ni2P (red), Ru–Ni2P (blue), and Ru (black). Images display the reactants, transition state, and product structures over Fe–Ni2P (top) and Ru–Ni2P (bottom).
Engproc 56 00053 g003
Table 1. The binding energies (ΔEads in kJ mol−1) of N2*, H*, and NHx (x = 0–3) over the catalytic surfaces examined in this study.
Table 1. The binding energies (ΔEads in kJ mol−1) of N2*, H*, and NHx (x = 0–3) over the catalytic surfaces examined in this study.
Ni2P(001)Fe–Ni2P(001)Ru–Ni2P(001)Fe(110)Ru(001)
SpeciesModeΔEadsModeΔEadsModeΔEadsModeΔEadsModeΔEads
N2*M1−25M1−35M1−50M1−95M1−34
H*M3−222M3−240M3−246M3−288M3−262
N*M3−375M3−435M3−445M4−601M3−551
NH*M3−325M3−353M3−365M3−528M3−437
NH2*M2−224M2−235M2−256M2−326M3−213
NH3*M1−60M1−50M1−74M1−109M1−63
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Almithn, A. Catalytic Performance of Doped Ni2P Surfaces for Ammonia Synthesis . Eng. Proc. 2023, 56, 53. https://doi.org/10.3390/ASEC2023-15319

AMA Style

Almithn A. Catalytic Performance of Doped Ni2P Surfaces for Ammonia Synthesis . Engineering Proceedings. 2023; 56(1):53. https://doi.org/10.3390/ASEC2023-15319

Chicago/Turabian Style

Almithn, Abdulrahman. 2023. "Catalytic Performance of Doped Ni2P Surfaces for Ammonia Synthesis " Engineering Proceedings 56, no. 1: 53. https://doi.org/10.3390/ASEC2023-15319

Article Metrics

Back to TopTop